Anda di halaman 1dari 8

Plastic and Polymer Technology (PAPT) Volume 2 Issue 1, March 2013

www.seipub.org/papt

Physical Characterization of Poly (vinyl


pyrrolidone) and Gelatin Blend Films Doped
with Magnesium Chloride
E. M. Abdelrazeka, H.M. Ragabb, M. Abdelaziza,c*
a

Physics Department, Faculty of Science, Mansoura University, Mansoura, Egypt

Physics Department, Faculty of Science, Al-Azhar University (Girls), Egypt

Natural Science Department, Community College of Riyadh, King Saud University, kingdom of Saudi Arabia

emabdelrazek@yahoo.com; a,c*mabdelaziz62@yahoo.com

Abstract
Polymeric films of pure poly (vinyl pyrrolidone) (PVP) and
PVP/gelatin blend (60/40) containing various amounts of
MgCl2 were prepared using a casting technique. The
structural and related physical properties of the prepared
films were studied using different techniques. The obtained
data revealed that the addition of gelatin and magnesium
chloride causes structural variation in the PVP network. The
X-Ray diffraction (XRD) patterns reveal that the amorphous
nature of the blend increases with the concentration of
magnesium chloride. Complex formation was confirmed by
XRD and FT-IR analysis. The DSC results indicate that the
addition of both gelatin and magnesium chloride to PVP
changes the thermal behavior, such as the glass transition
temperature and thermal stability. The analysis of the UVvisible optical absorption shows the decrease in the optical
band gap (Eg) as dopant concentration increase. The reduced
values of the optical gap improve their optical response,
which can be used as optical sensors. The direct current (DC)
electrical resistivity studies revealed a linear temperature
dependence of the hopping distance (Ro) for all doped films.
The conduction mechanism discussed based on the phononassisted charge carrier inter-polaron hopping model.
Moreover, the addition of both gelatin and MgCl2 gives rise
to improve the electrical properties of PVP film. The optical
and electrical results suggested the applicability of these
materials in optical and/or electrical sensors.

Keywords
Blends; XRD; FT-IR; DSC; Optical; Electrical Properties

Introduction
Polymer science has shown a tendency in the last
decade to create blends of different polymers rather
than to develop new polymers. Blending of polymers
is one of the simplest means to obtain a variety of
physical and chemical properties from the constituent
polymers. The gain in newer properties depends on

the degree of compatibility or miscibility of the


polymers at a molecular level. Generally, the polymer
polymer miscibility is due to some specific interactions
like dipoledipole forces, hydrogen bonding and
charge transfer complexes between the polymer
segments.
Water soluble polymers are important from industrial
view point. Polyvinyl pyrrolidone (PVP) has excellent
characteristics such as high-dielectric constant,
dissolubility, stability, compatibility and resistance
and large scale screen printing of PVP films at low cost
is feasible. PVP is highly soluble in polar solvents such
as alcohol, so it is preferable to avoid phase separation
in the reaction. Another advantage of using PVP is
that PVP can be thermally crosslinked and that makes
the composites have outstanding thermal stability and
high
mechanical
strength.
Furthermore,
the
amorphous structure of PVP also provides a low
scattering loss, which makes it an ideal polymer for
composite materials for optical application. PVP
thermally decompose before reaching its molten state.
This terminates the application of this polymer. PVP is
chosen as a matrix for the composites because of the
two important characteristics. One is that PVP has
good film-forming and adhesive behavior on many
solid substrates and its films exhibit good optical
quality and mechanical strength. Another is that the
pyrrolidone group of PVP prefers to complex with
many inorganic salts resulting in fine dispersion and
surface passivation of them. Blending PVP with a
potentially useful natural biopolymer such as gelatin
seems to be an interesting way of preparing a
polymeric composite. Gelatin not only maintains
inherent biological activities of PVP, but also gains a
lot of new properties and functions. Such a composite

www.seipub.org/papt

Plastic and Polymer Technology (PAPT) Volume 1 Issue 1, March 2013

reported improved thermal stability and absorbability


when used as matrices for magnesium chloride. As
aforementioned, several researchers have reported of
PVP and it`s blends with other polymers. However,
there is little research reported on the preparation and
physical characterization of PVP and gelatin blend
doped with metal halide.
The present work is carried out to investigate the
effect of both doping gelatin and magnesium chloride
on the structure of PVP using XRD, FT-IR and DSC
techniques. Theses composites obtained would offer
opportunities to explore their novel optical, thermal
and electronic properties..
Experimental work

Sample preparation
Poly (vinyl pyrrolidone) (PVP) from Aldrich chemical
co. Ltd England were used as received. The quantity of
PVP and gelatin (60/40) by weight to weight (wt %)
was added to doubly distilled water with stirring the
solution at 343K to complete dissolution. Required
quantity (0.0, 2, 6, 8,10,15 and 20 wt%) of MgCl2 was
also dissolved in doubly distilled water and added to
the polymeric solution with continuous stirring. Then
the solution was left for 24 hour to eliminate bubbles.
The solution was poured onto cleaned Petri dishes and
dried in oven at 333 K for 4 days to ensure removal of
the solvent traces. After drying, the films were peeled
from Petri dishes and kept in vacuum desiccators until
use. The thickness of the obtained films was in the
range of 80-90 m for the FT-IR measurement and
100150 m for other measurements.

Measuring techniques
The X-Ray diffraction (XRD) scans were obtained
using a Seimens type F diffractometer with CuKa
radiation and LiF monochromator. FT-IR absorption
spectra were carried out using the single beam Fourier
transform-infrared spectrometer (FT-IR-430, JASCO,
Japan). FT-IR spectra of the samples were obtained in
the spectral range of 4000-400 cm-1. The differential
scanning calorimetry (DSC) measurements were
performed by using a Shimadzu DSC-50 apparatus in
the temperature range 303873 K with a heating rate of
10 K/min. Ultraviolet and visible (UV-VIS) absorption
spectra were measured in the wavelength region of
200 - 900 nm using spectrophotometer (V-570
UV/VIS/NIR, JASCO, Japan). The direct current
electrical resistivity was measured using an auto-range
multimeter (Keithley 175) with an accuracy of 2%.

Results and discussion

X-Ray diffraction (XRD)


The measured XRD profiles of pure PVP, PVP/gelatin
blend, PVP/gelatin: MgCl2 films and MgCl2 are shown
in FIG. 1. There was a noticeable change in the
intensity of XRD peaks of the doped samples in
addition to appearance of additional peaks. The pure
PVP scan shows very broad diffraction peak around
2 = 22.3o, which confirms the amorphous nature of
the prepared polymer film that is conformity with
either reported in literature. When gelatin was added
into PVP, making PVP/gelatin blend, the intensity of
this peak increased and became sharper. An increase
in the intensity and decrease in the width of
PVP/gelatin blend are observed indicating a
semicrystalline structure, which leads to their good
compatibility. Similar result was reported by
Nagahama et al. For the addition of MgCl2 to the blend
matrix, there is a shift in the diffraction peak towards
lower diffraction angles confirmed a complex
formation. Moreover, there is a scattering peak at 2 =
8.5o was observed for doping levels 10 and 20 wt%.
This peak belongs neither to PVP nor PVP/gelatin or
MgCl2. This peak may indicate the appearance of new
crystalline phase of the blend matrix. Furthermore, the
peaks for 2 = 21.5o, 23 o, 29 o and 36o, pertaining to
MgCl2 disappeared in the complexes which indicate
the complete dissociation of the salt in the polymer
matrix. This observation confirms that complexation
has taken place in the amorphous phase. The decrease
in the intensities of the peaks on doping shows that
the decrease in the crystallinity and simultaneous
increase in the amorphousity of the complexed films.
This amorphous nature is responsible for greater ionic
diffusivity resulting in high ionic conductivity.

Fourier transforms infrared analysis


FIG. 2 shows the FT-IR spectra of PVP/gelatin blend
film as well as PVP/geatin: MgCl2 films in the range of
2000 - 400 cm-1. Both the spectrum of PVP, not shown,
and PVP/gelatin films have nearly similar
characteristic FTIR bands. Spectrum of the blend film
shows intense peaks at 1661, 1466 and 1285 cm-1
corresponding to C=O, C=C and C=N respectively. A
broad peak located at 918 cm-1 is due to the outer face
vibration oscillation of the hydroxyl group ( O-H). The
other peaks centered at 1425and 1285 cm-1 are assigned
to the inner face bending vibrations of the hydroxyl
group. The spectra of PVP/gelatin: MgCl2 films show
some difference as follows;

Plastic and Polymer Technology (PAPT) Volume 2 Issue 1, March 2013

www.seipub.org/papt

2) New absorption bands in between 700 to 900 cm-1


are observed. The new bands may be correlated
likewise to defects induced by the charge transfer
reaction between the polymer chain and the
dopant.

FIG.1 XRD PATTERNS OF PURE PVP, PVP/GELATIN BLEND,


PVP/GELATIN BLEND DOPED WITH (6, 10 AND20 WT%) OF
MGCL2 AND MGCL2 SALT

1) The increasing in the intensity and shifting


towards higher wavenumber of the absorption
peak at about 627 cm-1, assigned to Mg-O-Mg
stretching, corresponded to strong interaction
between polymer matrix and dopant.

FIG. 2 FT-IR ABSORBTION SPECTRA OF PURE PVP/GLETAIN


BLEND AND PVP/GLETAIN BLEND DOPED WITH MgCl2

3) Shift of peak from 1617 cm-1to 1661 cm-1, indicated


strong columbic interaction between PVP and Mg
ions. Some researchers have proposed that the
shifting of the C = O group in PVP can be
attributed to the charge of p- conjugation
associated with the amide group of PVP arising
from dissociation of PVP chains due to the
incorporation of other species. It is possible that
the interaction between the dopant and PVP leads
to the dissociations of the aggregated PVP chains,
resulting in the shifting of the C=O vibration band.
From the summarized spectra features listed
above, it can be seen that stronger molecular
interaction exists between polymers and dopant.

Differential scanning calorimetry


The thermal properties of pure PVP as well as
PVP/gelatin blend and their complexed films were
examined by DSC to estimate how the thermal
transitions of the prepared films were affected by the
different concentrations of MgCl2 as shown in FIG. 3.
For pure phase of PVP, FIG. 3, the curve of DSC shows
a major endothermic peak centered at about Tg= 345 K,
assigned to a glass transition process associated with
and an enthalpy of 542.35 J/g, and extending from
about 303-398 K. It is probable that this broad peak
includes another small one at 373 K due to the removal
of some water trapped in the polymeric network. In
the literature, several conflicting values of glass
transition temperature (Tg) of PVP were found,
ranging from 316 K at a scanning rate of 5 K/min to
450 K at a heating rate of 25 K/min. Several reasons are
usually suggested for such confliction of Tg values
including: presence of impurities, changes in specific
heat involved, inability to attain near-equilibrium
conditions during measurement and/or the faster rate
of temperature change compared to the change in
molecular rearrangement, along with the large
influences of sorbed moisture due to the hydroscopic
nature of the material. The minor endothermic peak at
about 483K in the thermograms of DSC for PVP may
be attributed to a solid solid transition of unspecific
nature. Kumar et al reported a similar peak at 479 K.
In addition to the endotherms at 345 and 483 K, the
PVP sample showed an endothermic peak at about
Td1= 665 K, which can be attributed to the thermal

www.seipub.org/papt

Plastic and Polymer Technology (PAPT) Volume 1 Issue 1, March 2013

chemical decomposition process associated with an


enthalpy of 16.94 J/g.
On the other hand, FIG. 3 shows also the DSC
thermograms for PVP/gelatin blend. One can observe
that the addition of gelatin to PVP led to a shift of the
peaks position for Tg and Td1 towards higher
temperatures at about 363K and 683 K respectively. It
is to be mentioned also that the peak at 483K for PVP
disappeared in the thermogram of the blend.
Furthermore, FIG. 3 depicts the DSC curves for the
PVP/gelatin doped with different concentrations of
MgCl2. It can be seen that the increase in doping

different structures of polymeric matrix. This result is


in agreement with that found in XRD analysis. As a
result from pervious discussion, the increase of both
MgCl2 and gelatin into the polymeric matrix improved
the thermal stability of the PVP.

UV/Vis optical absorption


FIG. 4 displays the UV-vis spectra of PVP/gelatin
blend and blend doped with different content of
MgCl2. The spectra of undoped and doped blend had
intense band at 208 nm, which may have been due to
the presence of chromophric group of gelatin. This
result is in agreement with pervious reported. In
addition the spectra exhibit a broad absorption band
in between 245 and 285 nm. This band was assigned to
the existence of carbonyl- groups associated with
ethylene un-saturation (C = C) C = O. Moreover,
the absorption spectra of the complexed films
indicated an increase in the absorption intensity of the
gelatin and shifting of the PVP broad band toward
higher wavelength. The increase in the absorption
intensity of gelatin and the shifting of the band of PVP
with increasing MgCl2 confirmed the complex
formation between the polymer and the dopant.
The optical band gap (Eg) values were obtained from
the absorption coefficient spectra according to the
well-known energy exponential relation:

h = B op (h E g

(1)

where Bop is the film`s constant, h is the photon


energy, = 2 .303 A is the absorption coefficient, A is
d

the absorbance, d is the film thickness and the


exponent m depends on the kind of the optical
transition that prevails, specifically, the n values are
1
2

3
2

, 2, and 3 for transitions directly

directly forbidden, indirected


forbidden, respectively.
FIG. 3 DSC SCANS OF PURE PVP, PVP/GLETAIN BLEND AND
PVP/GLETAIN BLEND DOPED WITH DIFFERENT
COMPOSITIONS OF MGCL2

concentration led to a shift of the Tg towards higher


temperature, along with remarkable gradual increase
of enthalpy values. The increase in the Tg values is due
to the occurrence of grafting interactions of Mg2+ with
amine group. Besides, the endothermic peak assigned
to thermal decomposition of PVP at 665 K splits into
two thermal decomposition temperatures (Td1 and Td2)
at about 665 and 783 K. These two peaks are probably
of thermal decomposition nature associated to

allowed,

allowed,
direct

Energy band gap values were obtained by


extrapolating the straight line plot of (h) versus h
as shown in FIG. 5. For the present study, the
absorption relation n = 2, obtained by the best fit of
eq. 1 indicated indirected allowed band transition.
The linear dependence of the band gap on the dopant
concentration is shown in FIG.6 with numerical
equation:

E g = 0 . 049 W + 5 . 31

(2)

Plastic and Polymer Technology (PAPT) Volume 2 Issue 1, March 2013

www.seipub.org/papt

photon energy (figure not shown). The doping level


dependence of Er is shown in FIG. 6. It is observed that
the Urbach tail for pure blend is less than that for the
doped ones, which increases with linear behavior as
MgCl2 content increase. The increase in Er is due to the
increase of the defects states in the polymeric matrix.
The energy sum (Et= Eg+ Er) can be taken to represent
the mobility gap of the charge carriers existing in the
conductivity specimens. The calculated values of Er
show decreasing with increasing dopant concentration
indicating reduction of the charge trapping centers.

FIG. 4 DOPING LEVEL DEPENDENCES OF THE OPTICAL


ENERGY GAP Eg (eV)

where W represents to the doping level. It is clear


that, the monotonic decreasing in the values of the
optical energy gap of PVP/gelatin blend with
increasing the MgCl2 content is observed. The change
of Eg attributed to the change of induced energy states
due to the change of intercalation mode.This
confirmed the complex formation between the
polymeric matrix and Mg- ions. On other wards,
increasing MgCl2 content may cause the localized
states to overlap and extended in the mobility gap.
This overlap may give us an evidence for decreasing
energy gap when MgCl2 content is increased in the
polymeric matrix.

FIG. 5 THE ABSOPTION SPECTRA OF PVP/GLETAIN BLEND


ANDPVP/GLETAIN DOPED WITH MGCL2

The Urbach tail was found to be related directly to


similar exponential tail for the density of states of the
band gap. The width of the Urbach tail is an indicator
of defect levels in the forbidden gap. The following
relation was used to calculate the width of the Urbach
tail:

h
= 0 exp
Er

(3)

where o is a constant and Er is the width of the tail of


the localized states due to the effect levels in the
forbidden gap. The values of Er are calculated from the
slope of the linear dependence of the natural
logarithm of the absorption coefficient versus the

FIG. 6 DOPING LEVEL DEPENDENCE OF THE OPTICAL


ENERGY GAP Eg (EV)

DC electrical resistivity
In polymeric films, the change of electrical resistivity
( ) with temperature is due to the segmental motion,

www.seipub.org/papt

Plastic and Polymer Technology (PAPT) Volume 1 Issue 1, March 2013

which results an increase in the free volume of the


system. This increase in free volume would facilitate
the motion of ionic charge. Similar behavior was
observed in a number of other films. The DC electrical
resistivity of the polymer doped with metal halide
mainly depends of the actual concentrations of the
metal ions in the polymeric matrix and their mobility.
The DC electrical resistivity values can be calculated
from the following equation:

RS
d

(4)

where R is the resistance and S is the surface area of


the electrode, and d is the thickness of the investigated
sample.
Due to the presence of polarons and/or bipolaron,
detected by the FT-IR analysis, the dc electrical
resistivity ( ) could be discussed on the basis of
Kuivalainen et al modified inter-polaron hopping
model. In this model the conduction mechanism could
be interpreted on the basis of phonon-assisted charge
carriers hopping between polaron and bipolaron
bound states in the polymer. Thus the electrical
resistivity is expressed as:

2
Ro2 ( y p + ybp )
2BRo

=
exp
2
2
y p ybp
Ae ( T )

kT

inter-chain distance. Taking yp = ypb for simplicity,


which is an acceptable approximation and using
equations (5) and (6), we can obtain the values of the
hopping distance Ro. A linear temperature dependence
of Ro for all doping levels is shown in FIG. 7. This
indicates that the concentration of thermally activated
polarons (acting as hopping sites for the charge
carriers) increases gradually as the temperature
increases. FIG.8 displays the dependence of hopping
distance Ro on doping levels at T = 370 K. It can be
seen that the hopping distance decreases with
increasing MgCl2. This may be related to a possible
increase in the number of conduction paths created
between the metal particles aggregates in the
polymeric matrix in addition to a decrease in the
width of the potential barriers within the conductivity
regions. Therefore, more charge carriers may be able
to hop by tunneling, resulting in the decrease in the
resistivity.

(5)

where k is Boltzmann,s constant, A1= 0.45; B1 = 1.39; yp


and ypb are the concentration of polarons and
1

bipolarons, respectively and

Ro = [3 /(4Cimp)]3

is

the typical separation between impurities whose


concentration is Cimp; =(// 2)1/3 is the average decay
length of a polaron and bipolaron wave function; and
// and are the decay lengths parallel and
perpendicular to the polymer chain, respectively.
According to the calculations of Bredas et al, polarons
and bipolarons induce defects of the same extension.
The electronic transition rate between polaron and
bipolaron states can be expressed as:

T
(T ) = o

300 K

11

(6)

where the prefactor, o = 1.21017 s-1, was estimated by


Kivelson. The order of magnitude of in the present
work was adjusted with the impurity concentration
Cimp, which actually was the fitting parameter. The
parameter
// =1.06 nm, while 0.22 nm, which
depends on the inter-chain resonance energy and the

FIG. 7 TEMPERATURE DEPENDENCE OF HOPPING DISTANCE


(RO) FOR VARIOUS DOPED FILMS

FIG. 8 DOPING LEVEL DEPENDENCE OF HOPPING DISTANCES


RO FOR VARIOUS DOPED FILMS AT T = 370 K

Plastic and Polymer Technology (PAPT) Volume 2 Issue 1, March 2013

The DC electrical resistivity () was measured in the


temperature range of 335-400 K for the PVP/gelatin
blend treated with different contents of MgCl2. As
Arrhenius relation the dependence of electrical
resistivity () has the form:

Ea
= o exp

kT

(7)

where o is the proportionality constant, Ea is the


thermal activation energy, and k is the Boltzmann
constant. FIG.9 shows the temperature dependence of
electrical resistivity of the complexes in the range 335400 K. It is observed that as the temperature increases,
the electric resistivity decreases for all complexes and
this behavior is in agreement with the theory
established by Armand et al. This can be rationalized
by recognized the free volume model. When
temperature is increased, the vibrational energy of a
segment is sufficient to push against the hydrostatic
pressure imposed by its neighboring atoms and create
a small

www.seipub.org/papt

The variation of logarithm of the electric resistivity ()


as a function of composition of MgCl2 in PVP/gelatin
blend at 370 K is shown in FIG. 10. It continued to
decrease with increasing dopant concentration. The
fast decrease in electric resistivity at lower dopant
concentration of MgCl2 is attributed to the formation
of charge transfer complexes or decrease in the
crystallinity of the blend, while the slow decrease at
higher dopant concentrations is due to the formation
of ionic aggregates.
The thermal activation energies were calculated from
the slope of these plots, and the values are shown in
FIG.10. These values are found to decrease with
increasing concentration of MgCl2. This may be due to
the fact that the addition of small amounts of dopant
forms charge transfer complexes in the host lattice.
These charge transfer complexes increase the electrical
conductivity by providing additional charges in the
lattice. This results in a decrease of activation energy.
As a result both the electrical resistivity ( ) and
thermal activation energy (Ea) behaves nearly
monotonically, this lead to important electrical
technological applications.

FIG.10 DOPING LEVEL DEPENDENCE OF HOPPING


DISTANCES RO FOR VARIOUS DOPED FILMS AT T = 370 K

Conclusion
FIG. 9 TEMPERATURE DEPENDENCEOF THE LOGARITHM OF
ELETRICAL RESISTVITY () FOR VARIOUS DOPED WITH
DIFFERENT CONTENTS OF OF MGCL2

amount of space surrounding its own volume in


which around the polymer chain causes the mobility
of ions and polymer segments and hence the electrical
resistivity. Hence, the increment of temperature causes
the decrease in electrical resistivity due to increased
free volume and their respective ionic and segmental
mobility.

PVP/gelatin blend and it`s complexed films were


prepared using a solvent casting technique. The XRD
study reveals the amorphous nature of the blends and
it`s complex films. Complex formation between the
polymeric matrix and Mg ions was confirmed by both
XRD and FT-IR analyses. DSC scans indicated that the
thermal stability of the PVP is improved when
blended with gelatin and loaded with MgCl2.
Incorporation of magnesium chloride in PVP/gelatin
blend increases the charge carriers, which causes a

www.seipub.org/papt

Plastic and Polymer Technology (PAPT) Volume 1 Issue 1, March 2013

decrease of both optical band gap and electrical


resistivity of the prepared films. The decrease in
optical band gap on doping levels was assigned to the
formation of the charge transfer complex in the host
lattice. The reduced values of the optical gaps improve
their optical response. These films can be used as
microwave sensors. The conductivity study indicates
that the polymer blend can be effectively doped with
MgCl2 to improve its conductivity. The increase in
conductivity is attributed to the formation of charge
transfer complexes. The calculate optical band gap
values of these materials as well as the magnitude of
their electrical conductivities of the thin films suggest
the possibility of considering them for use in the
preparation of electronic devices.
REFERENCES

Kaplan H. and Guner A., J. Appl. Polym. Sci. 78 (2000): 9941001.


Kivelson S., Phys. Rev. B 25 (1982) 3798.
Kuivalainen P., Stubb H., Isotalo H., Yli P., Holmstrom C.,
Phys. Rev. B 31 (1985): 7900.
Kumar V., Yang T. and Yang Y., Int. J. Pharm. 188 (1999): 221.
Latour M., Anis K., Faria R.M., J. Phys. D 22 (1989): 806.
Lokamatha K.M., Bharathi A., Kumar S.M.S. and Ramarao
N., Int. J. Pharm. Pharm. Sci., 2 (2010): 169.
Mott N.F. and Gurney R.W., `Electronic(1940) 34.
Muthyala S., Bhonde R.R. and P.D. Nair, Islets 26 (2010) 357.
Nagaahama H., Maeda H., Kashiki T., Jayakumar

R.,

Furuike T. and Tamura H., Carbohydrate Polym. 79


(2009): 255.
Rajendran S., Sivakumar M., Subadevi R., Mater. Lett., 58

Abd El-Kader F.H., Gafer S.A, Basha A. F., Bannan S.I. and
Basha M.A.F., J. Appl. Polym. Sci., 118 (2010): 413.
Abdelaziz M., J. Magn. Magn. Mater. 2792 (2004): 184. Basha
A.F. and Basha M.A.F., Polym. Bull. 68 (2012): 151.
Bredas J.L., Chance R.R., Silbey R., Phys. Rev. B 26 (1982):
5843.
Caykara T., Demirci S. and Kantoglu O., Polym. Plast.
Technol. Eng. 46 (2007): 737.
Chobagno M.B., Duclot J.M., Vashishta P., Mundy L.N. ,
and Shenoy G., Fast-Ion Transport in Solid, NohHolland, Amsterdam, (1979) 131.
Davis E.A. and Mott N.F., Philos. Mag. 22 (1970): 903.
Elasmawi I.S. and Adel Baieth H.E., Cur. Appl. Phys. 12
(2012) 141.
Elimat Z.M., J. Phys. D: Appl. Phys. 39 (2006) 2824.
Feng W., Tao H. and Liu Y., J. Mater. Sci. Technol., 22 (2006):
230.
Hatta F.F., Yahya M.Z.A. Ali A.M.M., Subban R.H.Y., Harun
M.K. and Mohamed A.A., Ionics 11(2005): 418.

(2004): 641.
Ravi M., Pavani Y., Kumar K.K., Bhvani S., Sharma A.K. and
Rao V.V.R. N., Mater. Chem. Phys. 130 (2011): 442.
Rosiak J., Burczak K., Olejniczak J. and Pekala W., Polimery
medycynie 88 (1989): 4-8.
Sengwa R. and Sankhla S., Coll. Polym. Sci. 285 (2007): 1237.
Sessa D.J., Woods K.K., Mohamed A.A. and Palmquist D.E.,
Ind. Crops Prod. 33 (2011): 57.
Sionkowska A., Kozlowska J. , Planecka A., Oanna J. and
Wisnwska S., Polym. Degrad. Stab. 33(2008): 1.
Sivaiah K., Rudramaddvi B.H. and Buddhudu S., Indian J.
Pure & Appl. Phys. 48 (2010): 658.
Sivaiah K. and Buddhudu S., Indian J. Pure &Appl. Phys., 49
(2011): 377.
Subba V., Han X., Zhu Q.Y. and Qiang L., Microelectr. Engin.
8 (2006): 281-288.
Umadevi C., Mohan K.R., Achari V.B.S., and Sharma A.K.,
Ionics 16 (2010): 751.
Urbach F., Phy. Rev., 92 (1953): 1324.

Anda mungkin juga menyukai